Next Article in Journal
Light-Activated Modified Arginine Carbon Dots as Antibacterial Particles
Next Article in Special Issue
Sodium Methoxide Catalysed One-Pot Glycidol Synthesis via Trans-Esterification between Glycerol and Dimethyl Carbonate
Previous Article in Journal
Photocatalytic Activity of ZnO/Ag Nanoparticles Fabricated by a Spray Pyrolysis Method with Different O2:N2 Carrier Gas Ratios and Ag Contents
Previous Article in Special Issue
Kinetics of the Direct DME Synthesis: State of the Art and Comprehensive Comparison of Semi-Mechanistic, Data-Based and Hybrid Modeling Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Perspective on CO2 Hydrogenation for Dimethyl Ether Economy

Key Laboratory of Syngas Conversion of Shaanxi Province, School of Chemistry and Chemical Engineering, Shaanxi Normal University, Xi’an 710119, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1375; https://doi.org/10.3390/catal12111375
Submission received: 14 October 2022 / Revised: 29 October 2022 / Accepted: 3 November 2022 / Published: 6 November 2022

Abstract

:
The CO2 hydrogenation to dimethyl ether (DME) is a potentially promising process for efficiently utilizing CO2 as a renewable and cheap carbon resource. Currently, the one-step heterogeneous catalytic conversion of CO2 to value-added chemicals exhibits higher efficiency than photocatalytic or electrocatalytic routes. However, typical catalysts for the one-step CO2 hydrogenation to DME still suffer from the deficient space–time yield and stability in industrial demonstrations/applications. In this perspective, the recent development of the one-step CO2 hydrogenation to DME is focused on different catalytic systems by examining the reported experimental results and the reaction mechanism including the catalytic nature of active sites, activation modes and of CO2 molecules under relevant conditions; surface intermediates are comparatively analyzed and discussed. In addition to the more traditional Cu-based, Pd-based, and oxide-derived bifunctional catalysts, a further emphasis is given to the characteristics of the recently emerged In2O3-based bifunctional catalysts for the one-step conversion of CO2 to DME. Moreover, GaN itself, as a bifunctional catalyst, shows over 90% DME selectivity and a reasonably high activity for one-step CO2 hydrogenation, and the direct hydrogenation of CO2 via the unique non-methanol intermediate mechanism is highlighted as an important illustration for exploring new catalytic systems. With these analyses and current understandings, the research directions in the aspects of catalysis and DME economy are suggested for the further development of one-step DME synthesis from CO2 hydrogenation.

1. Introduction

The increasingly severe issue caused by the cumulative over-emission of greenhouse gas (GHG) is one of the most prominent concerns in the world. Among commonly concerned GHGs, carbon dioxide (CO2), which is mostly derived from the unlimited utilization of fossil fuels, is believed to be a main contributor to the greenhouse effect, global warming and climate change [1,2,3]. Therefore, reducing the total emission of CO2 becomes an important solution to deal with possible climate change in future.
The capture, storage and utilization (CCSU) CO2 has been considered an effective method to alleviate the problem [4,5,6,7]. Chemical conversion of CO2 to high-value-added products exhibits dramatic efficiency in economic and industrial sectors, and has been paid increasing attention. Because the CO2 molecule exhibits high thermodynamic stability but can be employed as a weak oxidant [8,9], conversion of CO2 through hydrogenation becomes an ideal procedure [1,2,10]. Moreover, considering current techniques, the heterogeneous hydrogenation of CO2 should be a more promising technique due to its higher catalytic activity than photocatalytic or electrocatalytic techniques.
In the procedure of heterogeneous CO2 hydrogenation, various products can be obtained. These products can be divided into three types: (i) carbon monoxide (CO), formed from a reversed water–gas shift (RWGS) reaction; (ii) hydrocarbons (HCs), including methane, lower olefins, gasolines, diesels, aromatics, etc.; (iii) oxygenates, including methanol, dimethyl ether (DME), higher alcohols (C2+ alcohols), etc. Among them, CO is an intermediate product that needs to be further converted to other products via Fischer–Tropsch synthesis. Moreover, when atom efficiency is taken into account, one of the oxygen atoms in CO2 can remain during the CO2 hydrogenation to oxygenates, indicating that the atom efficiency for producing oxygenates is better than producing HCs.
In the case of the CO2 hydrogenation to oxygenates, single-carbon oxygenates, methanol and DME have been paid more attention due to their higher selectivity than C2+ alcohols. Previously, Olah et al. [11,12,13] proposed an industrial carbon cycle based on producing and converting methanol. They believed that in this process, the efficient utilization of CO2 could be realized by the thermochemical or electrochemical hydrogenation with the hydrogen generated from the electrolysis of water. As the intermediate product (Figure 1), methanol can be further converted into high-value-added chemicals, e.g., lower olefins and aromatics. Moreover, in spite of its toxicity, liquid methanol is feasible in the transportation sector; thus, it can also be employed as fuel for fuel cells and as substitute to gasoline due to its high octane rating.
However, DME can also be employed as an intermediate product of CO2 hydrogenation (Figure 1). During the process of CO2 hydrogenation, formation of DME exhibits a higher energetic efficiency than formation of HCs or alcohols [14], e.g., 97% of energy is stored in DME during CO2 hydrogenation. Moreover, DME is a non-toxic, non-carcinogenic and non-corrosive chemical applied as a propellant of cosmetic products. DME is also employed as a green refrigerant due to its low global warming potential, and as a promising ultra-clean fuel alternative to liquefied petroleum gas and diesel due to its high cetane rating [15]. More importantly, for both CO hydrogenation and CO2 hydrogenation, producing DME is thermodynamically more favorable than methanol. These advantages demonstrate that as the intermediate in the CO2 hydrogenation, DME exhibits more potential than methanol. Therefore, a new chemical economy centered on synthesizing and converting DME, the so-called DME economy [16], is expected to be significantly developed in the chemical industry in the future.
Thus, it is necessary and important to improve the technique of CO2 hydrogenation to DME. There are two kinds of processes to realize the CO2 hydrogenation to DME (Figure 1): (i) two-step process, in which CO2 is firstly hydrogenated to methanol in one reactor, and the methanol is further converted to DME in another reactor; (ii) one-step process, in which CO2 is hydrogenated to DME in the one and only reactor. However, irrespective of the one-step or two-step process [17,18], methanol is employed as the intermediate and two kinds of active sites are employed to catalyze the CO2 hydrogenation to methanol and, subsequently, methanol to DME. Thus, in the one-step CO2 hydrogenation to DME, these two active sites are coupled in one bifunctional catalyst, while the two-step CO2 hydrogenation to DME can be regarded as a special example of one-step CO2 hydrogenation to DME without any interaction between the two active sites. However, considering the total cost and complexity of the process, the one-step CO2 hydrogenation to DME is a more promising process in the DME economy.
According to current research [19,20,21,22], the design of an effective bifunctional catalyst is the most important challenge in one-step CO2 hydrogenation to DME, and a great deal of effort has been made. Over the bifunctional catalysts, metal-based catalysts are usually employed for CO2 hydrogenation to methanol, and the reduced Cu is the most studied active site. For the acidic sites catalyzing the dehydration of methanol to DME, different kinds of zeolites are the most employed under the atmosphere of CO2 hydrogenation. Based on these results, several reviews [17,19,23,24,25,26] have been published to provide conclusions and perspectives about CO2 hydrogenation to DME. Most of publications focus on the effects of promoters or supports on the catalytic performance. Moreover, the preparation method [20,23,24,27], the design of reactors [17,19,21,24,26] and thermodynamic considerations [18,19,23] are paid attention to, while deactivation and regeneration of catalysts are also involved [28]. To further reveal the effect on the catalytic performance, the nature of active sites, reaction mechanism and acidity of the acidic sites have been investigated [18,21,29,30]. However, in the previous publications, various competing conclusions are obtained and a general agreement needs to be reached.
Due to the advantage of the DME economy and the challenge in the design of an effective catalyst for CO2 hydrogenation to DME, this paper summarizes the recent developments in the investigation of one-step hydrogenation to DME. A thermodynamic approach is presented first. Then, we focus on the nature of active sites and the reaction mechanism over different types of catalysts, e.g., Cu-based catalysts, precious metal-based catalysts, oxide-based catalysts and GaN-based catalysts. According to the previous results, conclusions about the active nature and reaction mechanism over different catalysts are proposed. Based on these analyses, future perspectives are proposed.

2. Thermodynamic Study

In the process of CO2 hydrogenation to DME, methanol is usually present as an intermediate product. Moreover, CO is also formed as a principal by-product, and it can also be hydrogenated to methanol or DME. Therefore, thermodynamic properties of different reactions are studied, and results are listed in Table 1. It is observable that the RWGS reaction is endothermic, while the hydrogenation and dehydration reactions are exothermic. Consequently, the increasing reaction temperature significantly inhibits the DME selectivity and improves the CO selectivity, resulting in the limitation of the formation of DME. To achieve highly selective conversion of CO2 to DME, the operation temperature is usually kept low, e.g., <300 °C for the metallic catalysts [30] and <400 °C for the oxide- or GaN-based catalysts [31,32,33]. However, low reaction temperature also lowers catalytic activity in CO2 hydrogenation to DME. Therefore, improving the catalytic activity by reasonable design of the catalyst is an important challenge in CO2 hydrogenation to DME.
The thermodynamic results also demonstrate that the CO2-to-DME reaction exhibits lower enthalpy and lower Gibbs free energy than the CO2-to-methanol reaction. Shen et al. [35] calculated the equilibrium yield of DME, CO and methanol in the CO2 hydrogenation reaction (Figure 2) and found that under the same reaction conditions, the equilibrium yield of DME is much higher than that of methanol in the CO2 hydrogenation process. This result is consistent with other publications [34,36,37,38,39,40], and authors also indicate that the formation of DME can be improved by high pressure and low temperature. Based on the thermodynamic parameters, Catizzone et al. [41] studied the CO2 equilibrium conversion for CO2 hydrogenation, and found that CO2 hydrogenation to DME exhibited higher catalytic activity than CO2 hydrogenation to methanol in the temperature range of 160−320 °C and pressure range of 10−100 MPa. However, when the reaction temperature further increases (>350 °C), there is only small difference between the selectivity of DME and methanol. Moreover, when partial pressure of H2 becomes lower, CO2 hydrogenation to DME exhibits a higher catalytic activity than CO2 hydrogenation to methanol [34], indicating that the consumption of H2 can be lower for CO2 hydrogenation to DME. Therefore, these results demonstrate that CO2 hydrogenation to DME shows advantages in the thermodynamic parameters, compared with CO2 hydrogenation to methanol.
One-step CO2 hydrogenation to DME should be more efficient than two-step CO2 hydrogenation to DME. In the case of the one-step procedure, the thermodynamic results are the same as those of the CO2-to-DME reaction shown in Figure 2. In the case of the two-step procedure, we can assume that the equilibrium conversion is achieved in each reactor, then the final equilibrium thermodynamic results can be calculated. The thermodynamic results of methanol dehydration to DME are shown in Figure 3a. We observe that in the methanol dehydration [42], the methanol conversion slightly decreases with increasing reaction temperature, while the DME selectivity remains at nearly 100%. Further, the equilibrium DME yields for the one-step and two-step CO2 hydrogenation to DME are calculated, shown in Figure 3b. It is observable that the DME equilibrium yield in the one-step procedure is higher than that in the two-step procedure, indicating that the one-step CO2 hydrogenation to DME is thermodynamically favorable. Therefore, it could be a promising technique in the future DME economy.

3. Copper-Based Catalysts

There is no doubt that Cu-based catalysts are the most studied in previous works. Cu/ZnO/Al2O3 (CZA) coupling with acidic sites, a well-known bifunctional catalyst, is firstly applied in CO hydrogenation to DME [43]. Subsequently, its activity in CO2 hydrogenation to DME has been found and a great deal of study has been conducted. However, there is still no industrial application of Cu-based catalysts in CO2 hydrogenation to DME, because the catalytic performance, e.g., DME selectivity, catalytic activity and stability, needs to be improved. To achieve this, investigating the nature of active sites and the reaction mechanism becomes necessary.

3.1. Active Sites

Previous work has already demonstrated that over the Cu-based catalyst, methanol is firstly formed on the Cu-based metallic sites, and DME is subsequently formed via dehydration of methanol on the acidic sites. In this case, the Cu sites play an important role in activating CO2 and H2. This is the initial step of CO2 hydrogenation to DME and significantly affects the catalytic performance. In spite of its important role, the exact nature of Cu sites is still unclear. In an early study [44,45], the Cu-based catalyst becomes more active after the reduction in hydrogen flow. Thus, the metallic Cu (Cu0) species are speculated to be the active sites for CO2 hydrogenation. Further, temperature-programmed desorption of N2O (N2O-TPD) is developed to exactly characterize the surface area of Cu0 species, and the catalytic activity of CO2 hydrogenation to methanol is found to linearly increase with increasing Cu0 surface area [44,46,47,48,49,50,51]. Detailed results are shown in Figure 4a. Therefore, the researchers can declare that the Cu0 species are responsible for the catalytic activity in CO2 hydrogenation.
However, other researchers reported totally different results. Some publications [52,53,54,55,56,57,58] proposed that no reasonable relation could be shared between the Cu0 surface area and the catalytic activity. Moreover, some authors even found that a too high Cu0 surface area would inhibit the selectivity of methanol. The work of K. Stangeland et al. [59] indicated that the formation of methanol should not be attributed to the increasing Cu0 surface area (Figure 4b), while the formation of CO was improved by the increasing Cu0 surface area. Therefore, it is still doubtable that the Cu0 species are the only active sites in CO2 hydrogenation. These results made the researchers focus on other possibilities of the active nature over the Cu-based catalyst.
During CO2 hydrogenation to DME, Cu0 species are not the only detectable Cu-based species on the catalyst surface. Frei et al. [60] reported that Cu+ species are also detected during the reduction of Cu-based catalysts, although most of them turned to Cu0 species after the reduction. This conclusion is agreed with in the work of Natesakhawat et al. [47] and Zhou et al. [48] However, other researchers propose [53,61,62,63] different results, e.g., the Cu+ species can be found during and after the reaction of CO2 hydrogenation. In this case, researchers started to pay attention to the role of Cu+ species over the Cu-based catalysts in the CO2 hydrogenation. Din et al. [53] confirmed the presence of Cu+ species, and they also reported that the catalytic activity was dependent on the amount of neither Cu0 nor Cu+ species. Tan et al. [46] reported that although the catalytic activity increased with increasing Cu0 surface area, no simple relation could be obtained when the Cu+ species were involved. Furthermore, Sun et al. [61] detected both Cu0 and Cu+ species by Cu Auger spectra. They found that increasing catalytic activity was dependent on the amount of either Cu0 or Cu+ species; however, correlation of the catalytic activity with the amount of Cu+ species was better than that with Cu0 species. The results of Sun et al. [61], Dasireddy et al. [62] and Stangeland et al. [63] proposed that the enhanced catalytic activity was attributed to the total exposed Cu surface, and a higher molar ratio of Cu+/Cu0 species could also improve the catalytic activity (Figure 4c). Based on the notion that the Cu+ species may also be active in CO2 hydrogenation, Zhang et al. [64] and Pu et al. [65] employed Cu2O as the active phase and prepared Cu2O/ZnO and Cu2O/CeO2 catalysts, respectively. Their results demonstrate that the Cu+ species can catalyze CO2 hydrogenation to methanol; however, the catalytic activity of Cu2O supporting catalysts is lower than the typical CZA catalyst.
Figure 4. Correlation of space–time yield of methanol (STYmethanol) with (a,b) Cu surface area (S.A.) and (c) molar ratio of Cu+/Cu0. The Cu S.A. is measured by temperature-programmed desorption of N2O. The ratio of Cu+/Cu0 is characterized by X-ray photoelectron spectroscopy. Reaction conditions: (a) T = 240 °C, P = 3 MPa, GHSV = 3000 mL·g−1·h−1, H2/CO2 = 3. (b) T = 230 °C, P = 3 MPa, GHSV = 3800 mL·g−1·h−1, H2/CO2 = 3. (c) T = 200 °C, P = 2 MPa, GHSV = 2000 mL·g−1·h−1, H2/CO2 = 3. Reproduced with permission from Natesakhawat et al. [47], Stangeland et al. [59] and Dasireddy et al. [62], respectively published by ACS Publications, 2012, Elsevier, 2021, and Elsevier, 2019.
Figure 4. Correlation of space–time yield of methanol (STYmethanol) with (a,b) Cu surface area (S.A.) and (c) molar ratio of Cu+/Cu0. The Cu S.A. is measured by temperature-programmed desorption of N2O. The ratio of Cu+/Cu0 is characterized by X-ray photoelectron spectroscopy. Reaction conditions: (a) T = 240 °C, P = 3 MPa, GHSV = 3000 mL·g−1·h−1, H2/CO2 = 3. (b) T = 230 °C, P = 3 MPa, GHSV = 3800 mL·g−1·h−1, H2/CO2 = 3. (c) T = 200 °C, P = 2 MPa, GHSV = 2000 mL·g−1·h−1, H2/CO2 = 3. Reproduced with permission from Natesakhawat et al. [47], Stangeland et al. [59] and Dasireddy et al. [62], respectively published by ACS Publications, 2012, Elsevier, 2021, and Elsevier, 2019.
Catalysts 12 01375 g004
As the catalytic activity of CO2 hydrogenation is attributed to both the Cu0 and Cu+ species, researchers need to discuss the active nature of Cu-based catalysts considering both of them. Some researchers [55,66,67] confirmed the co-existence of Cu0 and Cu+ species on the catalyst surface, and the increasing amount of both Cu0 and Cu+ species favored the catalytic activity. Moreover, recent work [62,68] indicated that in CO2 hydrogenation to methanol, the increasing CO2 conversion was caused by a high number of Cu0 species, while a high number of Cu+ species improved the selectivity of methanol. Therefore, it is more acceptable that both Cu0 and Cu+ species are active in the CO2 hydrogenation, and these species can be labeled as Cuδ+ with 0 ≤ δ < 2.
Furthermore, researchers also focus on the effect of ZnO on the active nature and catalytic performance over the CZA catalyst, because ZnO is a necessary and important promoter in Cu-based catalysts. Results [69,70,71,72,73] of operando infrared spectroscopy (operando-IR) and density functional theory (DFT) calculations showed that CO2 hydrogenation over the ZnO-supported Cuδ+ sites or the Cu/ZnO interface is more favorable than bulk Cuδ+ sites. This is because the activation energy for the formation of format intermediates is lowered due to the presence of ZnO. Moreover, ZnO is also found to improve the reducibility of Cu oxides [44,60], improve the dispersion of Cuδ+ species [44,69,73] and enhance the adsorption of CO2 [70,72]. Some work [49,51] also proposed that the addition of Al2O3 enhanced the electron density of ZnO species, resulting in a stronger CO2 adsorption ability. Although the Cu–ZnO interface and Cu–Zn alloy are also considered active sites in some publications [69,71], most of the results agree that the reaction mechanism is the same over Cuδ+ with or without ZnO. Therefore, the Cuδ+ species should be considered essential active sites for Cu-based catalysts in CO2 hydrogenation.
In all, although the reduced Cu species have been confirmed as active sites, more efforts still need to be taken in future to further understand the exact function of Cu0 and Cu+ species. The effects of other important promoters that significantly improve the catalytic activity, e.g., ZrO2 [44] and Ga2O3 [74], also need to be investigated. These efforts can raise our awareness on the active nature of Cu-based catalysts and are important for designing catalysts with high performance.

3.2. Reaction Mechanism

The reaction mechanism is another important topic for Cu-based catalysts. Different from the active nature, the researchers have agreed that there are two mechanisms in Cu-catalyzed CO2 hydrogenation, the mechanism via format (HCOO*) intermediate and the other via CO* intermediate. Additionally, the mechanism via carbonyl (COOH*) intermediate is also proposed in a previous work [75]. However, the COOH* intermediate can easily turn into CO* and is further transformed to methanol. This indicates that in the mechanisms via COOH*, CO* is also a key intermediate. The mechanisms via COOH* and CO* are essentially the same.
The mechanism via HCOO* intermediate is considered the most important pathway in the Cu-catalyzed CO2 hydrogenation to methanol [75,76,77,78,79,80]. This mechanism is confirmed by operando-IR [58,75,77,78,81], 14CO2 [80]/D2 [75,78] isotope experiments, kinetic study [79] and DFT calculations [58,76]. As shown in Figure 5, CO2 is firstly hydrogenated to form HCOO*. Subsequently, one O atom in HCOO* is hydrogenolyzed and HCOO* is further hydrogenated to methoxyl (CH3O*) intermediate. Finally, the methanol is formed through the hydrogenation of CH3O*. Furthermore, HCOO* is found to be transformed to both methanol and CO.
However, the mechanism via CO* intermediate cannot be totally ignored. Liu et al. [82] reported that both CO* and HCOO* were formed during CO2 hydrogenation, but further conversion of HCOO* to methanol was more difficult than CO*. Therefore, they believed that the reaction mechanism of CO2–CO–methanol was more reasonable. Moreover, some results of DFT calculations [71,83,84] and kinetic study [83] also confirm the mechanism via CO*. Thus, researchers also studied the co-existence of the two mechanisms. Yu et al. [85] found that the Cu0 species catalyzed the formation of methanol via HCOO*, while the RWGS reaction also occurred. Subsequently, the formed CO could be transferred to the Cu+ species and be further hydrogenated to methanol. Their results indicated that both the CO2-to-methanol and RWGS reactions were catalyzed by the Cu0 species, and this is also confirmed by Kunkes et al. [86] Moreover, Poto et al. [87] reported that the reaction conditions could affect the reaction mechanism. Particularly, the mechanism via CO* was more favorable with high reaction temperature and high H2/CO2 ratio. With decreasing temperature and stoichiometric ratio of H2/CO2, the mechanism via HCOO* became more important. These results can explain why in a previous study under the typical conditions (<240 °C, H2/CO2 = 3) for CO2 hydrogenation to methanol, the mechanism via HCOO* was speculated to be principal.

3.3. Stability

Cu-based catalysts often suffer from poor stability. Three reasons cause this problem [28]: (i) the metallic Cu can be easily sintered at quite a low temperature of 240 °C; (ii) the coke formation can deactivate the metallic and acidic sites; (iii) the formed H2O in the reaction may oxidize the Cu0 species and deactivate the acidic sites of zeolites. To overcome this problem, ZrO2 is added as a promoter for the CZA catalysts [88], in order to slow the oxidation of Cuδ+ species. Modification of acidic sites is another important way to improve the stability [89]. Moreover, techniques of regeneration are also employed in the CO2 hydrogenation procedure to prolong the lifetime of catalysts. The corresponding results are summarized by Sobczak et al. [28].
However, in CO2 hydrogenation, more water is formed than in CO hydrogenation due to the RWGS reaction. Especially at increasing reaction temperature, the RWGS reaction is improved and even more water is formed, resulting in a serious deactivation of Cu-based catalysts. High reaction temperature also leads to sintering of the reduced Cu species. Therefore, to improve the catalytic activity together with the stability in one-step CO2 hydrogenation to DME, it is necessary to design a more stable active phase for CO2 hydrogenation to DME at high reaction temperatures.

4. Precious Metal Catalysts

Precious metals have been widely employed in heterogeneous catalysis due to their enriched unoccupied d-bands favoring the activation of small molecules such as CO, CO2, H2, etc. In the CO2 hydrogenation reaction, palladium is found to be active for hydrogenating CO2 to methanol. Thus, one-step CO2 hydrogenation to DME can be realized by coupling a Pd-based active phase and the acidic phase.
Most of the related publications [90,91,92,93] agree that the reduced Pd species (Pd0) are active sites for CO2 hydrogenation. However, agreement cannot be reached in the case of reaction mechanism. Bonivardi et al. [94] reported that HCOO* was the key intermediate over the Pd/Ga2O3 catalyst, and they also proposed a possible reaction pathway of HCO3 * (or CO3 *) → HCOO* → HCOOH* → CH3O* → CH3OH*. Moreover, the works of Schild et al. [95] and Arunajatesan et al. [96] also confirm the HCOO* mechanism. On the contrary, Malik et al. [97] proposed that over CaO–Pd/ZnO, CO2 was hydrogenated via the CO* intermediate, and similar results were obtained over the Pd/SiO2 catalyst in the work of Cabilla et al. [98].
Based on the current results, we can speculate that this difference is caused by the nature of active sites. As presented in Figure 6, on one hand, over the Pd-based catalysts with the oxides favorably activating the CO2 molecular, e.g., Ga2O3 [94], Al2O3 [96] or ZrO2 [95], the CO2 molecule is activated by the oxides but not by metallic Pd, according to the operando-IR results. In this case, HCOO* is characterized as the key intermediate. A typical case is proposed by Cardona-Martinez et al. [91], e.g., over the Pd/Ga2O3 catalyst, the metallic Pd sites are responsible for activating H2 and Ga2O3 is responsible for activating CO2. Subsequently, the activated CO2 can be transferred to the Pd surface and be hydrogenated to methanol. On the other hand, when more inert oxides, e.g., SiO2 [98] or ZnO [97], are employed into Pd-based catalysts, CO2 is adsorbed and activated by Pd0 species and the RWGS mechanism is dominated, indicating that on the active sites of Pd0 species, CO2 hydrogenation occurs via the CO* intermediate. Therefore, if only Pd is involved as the active site, CO2 is firstly hydrogenated to CO then further converted to methanol. This is different from Cu- and oxide-based catalysts with formate mechanisms.
In Table 2, typical examples of Cu- and Pd-based catalysts with high performances in CO2 hydrogenation to DME are presented. In the case of Cu-based catalysts, it is observable that different promoters significantly affect the catalytic activity and DME selectivity. The addition of La2O3 into the CZA/HZSM-5 catalyst clearly enhanced the formation of DME [99]. The utilization of ZrO2 [100,101] or Ga2O3 [74] to replace Al2O3 in the CAZ catalyst improves the space–time yield of DME (STYDME); however, the effect of CeO2 is less significant [102]. Moreover, the catalysts with ferrierite as their acidic sites exhibit better catalytic performance than those with HZSM-5 [100,103]. In the case of the Pt-based catalysts, the previous study indicates that addition of basic promoters such as CaO [90,98] and oxide supports such as Ga2O3 [91,94] or CeO2 [97] can improve the catalytic activity, because the adsorption and activation of CO2 can be enhanced. Some other works [104,105] also reported catalysts with a bimetallic active phase of Cu–Pd. However, the mixture of two kinds of active metal causes difficulties in understanding the active nature of the catalysts. Moreover, compared with the mono-metal-based catalysts (Table 2), the Cu–Pd bimetallic catalysts exhibit no significant enhancement in the catalytic performance.
It is also observable in Table 2 that the catalytic performance is clearly affected by the reaction conditions. Irrespective of the composition of catalysts, STYDME is improved by the increasing reaction pressure. Higher GHSV could enhance the DME selectivity and inhibit the CO2 conversion for Cu-based [100,101] and Pd-based catalysts [90]. However, the reaction temperature exhibits different effects on different catalysts. Over the CZA/HZSM-5 catalysts, the increase in the reaction temperature leads to a decreasing DME selectivity and a decreasing STYDME [106]. However, different from Cu-based catalysts, at a higher reaction temperature of 325 °C, the Ca-Pd/CeO2 catalyst exhibits a higher STYDME [90]. Moreover, R. Chu et al. [93] reported that the Pd/HZSM-5 catalyst exhibited a much better sulfur tolerance than the CZA/HSAM-5 catalyst (Figure 7).

5. Indium- and Gallium-Based Catalysts

5.1. Indium-Based Catalysts

In2O3 is found to be active in CO2 hydrogenation to methanol and a methanol selectivity of 80% is detectable even at a reaction temperature of >300 °C [109,110]. Researchers propose that the oxygen vacancies on defective In2O3 are active sites for CO2 hydrogenation. The DFT calculations [110,111] demonstrate that CO2 is activated over In2O3 via oxidation of the oxygen vacancies. After CO2 hydrogenation, the oxygen vacancies are reproduced by the reduction of H2 or CO. Moreover, results also confirm that the HCOO* species are the intermediate for CO2 hydrogenation over In2O3 catalysts.
Based on these facts, Pechenkin et al. [31] prepared a bifunctional catalyst of In2O3/halloysite nanotubes to catalyze CO2 hydrogenation. Their results (Figure 8) indicate that the DME selectivity reaches 65% at 200 °C; however, the increasing reaction temperature inhibits DME selectivity and improves methanol selectivity. At 300 °C, DME selectivity is only 20%. The authors declared that the exothermicity of the methanol dehydration inhibited the formation of DME with increasing reaction temperature. Therefore, in spite of the high stability, the catalytic activity and DME selectivity over the oxide-based catalysts are still low with increasing reaction temperature.

5.2. Gallium-Based Catalysts

As presented before in the work of Collins et al. [94], Ga2O3 is employed as the active site for activating CO2 over Pd/Ga2O3. Moreover, their results of operando-IR indicate that the methanol is formed via the format intermediate, the same as the Cu-based catalysts. It is worth noting that the presence of Pd in the catalyst does not change the reaction mechanism; it only improves the activation of H2 and the catalytic activity. Therefore, Ga2O3 can independently play a role in the activation and hydrogenation of CO2. Moreover, based on the results of the theoretical simulation, Studt et al. [112] prepare a Ni/Ga2O3/SiO2 catalyst for CO2 hydrogenation to methanol. Results indicate that the Ni/Ga2O3/SiO2 catalyst with Ni/Ga = 5/3 exhibits a similar yield of methanol as the CZA catalyst. This is because the theoretical simulation indicates that oxygen adsorption energy is closed over Ni/Ga2O3/SiO2 and CZA catalysts. The authors propose that the oxygen adsorption energy is related to the adsorption and activation of CO2 and further affects the selectivity. The addition of Ga2O3 enhances the oxygen adsorption energy as well as the methanol selectivity, due to the increasing ability of CO2 adsorption.
Therefore, the two above examples indicate the unique propriety of Ga2O3 for the adsorption and activation of CO2. Further, another compound of gallium, GaN, is also employed as an active phase for CO2 hydrogenation in the work of Liu et al. [33] The GaN catalyst exhibits a high DME selectivity under the conditions of T = 360 °C, P = 2.0 MPa, GHSV = 3000 mL·g−1·h−1, and H2/CO2 = 2. Furthermore, with decreasing contact time, the DME selectivity increases while the methanol selectivity decreases, indicating that in CO2 hydrogenation over GaN, DME is the primary product and methanol is the secondary product. The results of surface reaction further confirm DME as the primary product, while CO is also speculated as another primary product. As shown in Figure 9, the DFT calculations indicate that CO2 can be activated by different modes on the different GaN crystalline planes of (100) and (110). The following hydrogenation of CO2 leads to the formation of CH3 * and HCOO* intermediates, and the final formation of DME occurs on the interface of (100)/(110). Therefore, the reaction pathway over GaN is quite different from the metallic catalyst, indicating that GaN could be a totally new kind of active phase for CO2 hydrogenation to DME.

6. Transition Metal Oxide Catalysts

Different from the metallic species, metal oxides should not be sintered even at high temperatures of around 400 °C. Thus, enhanced stability can be expected over the metal oxide-based catalysts. Liu et al. [32] proposed a ZnAlOx catalyst prepared by the co-precipitation method (Figure 10). Under the reaction conditions of T = 320 °C, P = 3 MPa and GHSV = 12,000 mL·g−1·h−1, the DME selectivity in the products without CO can reach ca. 50% with a CO2 conversion of 5.2%. The active phase is speculated to be the oxygen vacancies over the crystallite of ZnAl2O4. Moreover, the results of in-situ IR spectroscopy demonstrate that CO2 is hydrogenated to methanol via the HCOO* and CH3O* species over the oxygen vacancies. Other authors [113] reported that the –OH groups on the surface of ZnAlOx also play an important role in CO2 hydrogenation.
However, prepared by the same method of co-precipitation and evaluated under a higher reaction temperature of 370 °C and a slightly lower GHSV of 10,800 mL·g−1·h−1, the ZnAl2O4 catalyst exhibits a DME selectivity of only 15% in the work of Wang et al. [113] Moreover, Song et al. [114] reported that the ZnAl2O4 catalyst, also prepared by the co-precipitation method, exhibited a methanol selectivity of >95% and a DME selectivity of <2%, under the conditions of T = 280 °C, P = 3 MPa and GHSV = 1800 mL·g−1·h−1. Therefore, it is uncertain whether ZnAl2O4 is an effective catalyst for the CO2 hydrogenation to DME, although it exhibits good stability in all the mentioned publications. It is necessary to find out the reason why the selectivity is so different regarding fact that the preparation method is the same and the reaction conditions are similar.

7. Acidic Sites and the Synergetic Effect

In the bifunctional catalysts, acidic sites play a role in converting methanol to DME (Figure 11). Thus, the properties of the acidic sites and the coupling of metallic and acidic sites significantly affect the catalytic activity, the DME selectivity and the stability. In a CAZ/acids catalyst [115,116], γ-Al2O3 is employed as an acidic site for CO hydrogenation to DME and a high selectivity can be obtained. However, for CO2 hydrogenation to DME, γ-Al2O3 is no longer a feasible acidic site. This is because the increasing accumulation of water in the CO2 hydrogenation, due to the RWGS reaction, leads to a serious deactivation of γ-Al2O3 with low resistance to water [27]. Therefore, the hydrophobic zeolites become the most used acidic sites in one-step CO2 hydrogenation to DME due to their good resistance to water.
Different kinds of zeolites have been employed as acidic sites, e.g., ZSM-5 [117,118], SAPO-11 [119], mordenite [120], ferrierite [100], etc. Recently, researchers focused on the effect of composition, textural properties and acidity on the catalytic performance. Frusteri et al. [121] reported that the ratio of Si/Al affected the selectivity of products and the stability through changes in the amount of acid and the resistance to water. A high Si/Al ratio enhances the resistance to water and the stability, but too high a Si/Al significantly lowers the surface acidity and catalytic activity. A low Si/Al ratio leads to high activity due to the increasing acidity, but the catalytic stability becomes poor. Further, by adjusting the Si/Al ratio from 27 to 127, the authors also found that over the CZA/HZSM-5 catalyst, the zeolite with Si/Al = 38 was optimized for stable hydrogenation of CO2 to DME with high DME selectivity.
There is no doubt that the acidity of zeolites plays a key role in the catalytic activity and DME selectivity. There are several parameters to characterize the acidity of zeolites [122,123,124,125,126], e.g., the amount of weak/strong acidic sites, the strength of acidic sites and the type (Bronsted or Lewis) of acidic sites. Typically, temperature-programmed desorption of NH3 is employed to characterize the amount and strength of acidic sites, and the in-situ IR spectroscopy of pyridine adsorption is employed to characterize Bronsted and Lewis acidic sites. In the case of the acidic type, both Bronsted and Lewis acidic sites are active sites for the dehydration of methanol to DME [27]; however, the Bronsted acidic sites also favor the dehydration of methanol to HCs [127]. Moreover, the Bronsted acidic sites also cause the formation of carbon residuals, resulting in the deactivation of catalysts. Regarding the strength, the increasing strength of acids can improve the catalytic acidity; however, acidic sites that are too strong catalyze the secondary conversion of DME and inhibit DME selectivity. Regarding the amount, an increase in the amount of weak Lewis acidic sites is found to be favorable in enhancing the catalytic performance in CO2 hydrogenation to DME.
Moreover, the porosity of zeolites also affects the catalytic performance [128,129]. The increasing distribution of mesopores with diameters of 10–50 nm is reported to favor the stability by inhibiting the formation of carbon residuals. However, the increasing distribution of micropores with diameters of smaller than 2 nm leads to increased formation of carbon residuals, resulting in deactivating the zeolites, although more micropores can improve the surface area of zeolites and the catalytic activity.

8. Conclusions and Research Directions

To conclude, one-step CO2 hydrogenation to DME is a promising process for the efficient utilization of CO2. However, the reasonable design of efficient catalysts is still a challenge. This paper summarized the recent process in heterogeneous catalysis for CO2 hydrogenation to DME, and focused on the nature of active sites and the reaction mechanism over different kinds of catalysts. In the case of Cu-based catalysts, it is generally agreed that the reduced metallic Cu species are the active sites, although the function of Cu+, Cu0 and Cuδ+ species is still unsure. The formate mechanism is dominated over the Cu-based catalysts, while the RWGS mechanism can be improved with increasing reaction temperature and increasing partial pressure of H2. The presence of ZnO promoter improves the adsorption and activation of CO2, thus favoring the catalytic activity. However, ZnO should not affect the reaction mechanism in CO2 hydrogenation to DME. Precious metal-based catalysts can also catalyze CO2 hydrogenation to DME. Metallic Pd and the Pd–Cu alloy are the most used active phases. Different from the Cu-based catalysts, The RWGS mechanism is dominated over the metallic Pd sites. However, with reference to the high cost, the catalytic performance over Pd-based catalysts is not significantly enhanced. The In2O3- and ZnAl2O4-based catalysts exhibit an excellent stability at high reaction temperatures. The oxygen vacancies are reported as the active sites and the formate species are speculated to be key intermediates. However, the catalytic activity and DME selectivity need to be improved. Over the GaN-based catalyst, DME is formed as a primary product while methanol is a secondary product. This is totally different from the metallic or oxide-based catalysts. The study of reaction mechanism indicates that both HCOO* and COOH* are key intermediates during the formation of DME. Based on current understanding, perspectives on the important research directions are proposed as follows:
(i)
Because of the RWGS reaction, CO is always formed as a by-product during CO2 hydrogenation, which lowers the selectivity of the high-value-added target products. Therefore, a deep understanding on the activation modes and subsequent transformation paths of CO2 molecules is very important for the rational design of a catalyst with limited CO selectivity. In the case of selective CO2 hydrogenation to DME, integrating the rich and available knowledge of the elementary steps into finding/re-optimizing the component and structure of catalysts via emerging techniques such as artificial intelligence is promising for inhibiting side reactions, including RWGS, which can be an important direction for the advancement of high-performance catalysts;
(ii)
In view of the exothermal nature of CO2 hydrogenation to DME and the easy sintering of metallic Cu, more stable Cu-based catalysts for the selective synthesis of DME are expectable at lower reaction temperatures, which brings on a great challenge for improving the catalytic activity. Indeed, the active site of Cu-based catalysts for both CO and CO2 hydrogenation reactions is still not unambiguously resolved. In this case, taking the redox nature of CO2 hydrogenation reactions into account, the disclosure of the dynamically evolved oxidation number of Cu and possible redox cycles between Cu0 and Cu+/Cuδ+ under the reaction conditions is crucial in identifying the active site. Thus, keeping these notions in mind, molecular-level understandings of CO2 hydrogenation to DME are reasonably expected with the development/modification of in-situ/operando techniques, which can be the second direction of the themed topic;
(iii)
In spite of their enhanced stability, catalytic activity and DME selectivity need to be improved for oxide-based catalysts. Previous works have consistently revealed the important role of oxygen vacancies. However, no simple correlation can be made between the amount of oxygen vacancies and the catalytic activity/DME selectivity. Thus, mechanistic studies on the formation/consumption of oxygen vacancies and the kinetics of the oxygen vacancies participating in elementary steps during CO2 hydrogenation are necessary works for the rational design and advancement of oxide-based catalysts for CO2 hydrogenation to DME with improved catalytic performance;
(iv)
Our results indicate that GaN itself is an effective bifunctional catalyst for the selective synthesis of DME from CO2 hydrogenation. More importantly, the unique mechanism, i.e., the direct hydrogenation of CO2 to DME without the formation of the methanol intermediate, is consolidated by the results of the catalytic experiments, operando spectroscopy, and DFT calculations. Indeed, the large-scale application of GaN as the catalyst is inevitably limited by the expensive Ga source. However, with the understanding of the wurtzite structure and the formation of DME over GaN cooperated by the two facets of (100), (110) and their interfaces, the mechanistically guided exploration of this novel, cheap bifunctional catalyst is attractive, which is worthy of being another research direction;
(v)
The summarized catalytic results indicate that one-step synthesis of DME via CO2 hydrogenation is a very promising candidate for CCSU. Moreover, the physical and chemical properties of DME in comparison with those of methanol make an attractive DME economy either parallel with or supplementary to the methanol economy. In both cases, however, investigations on both the expanded application of DME and the extension of DME-derived high-value-added chemicals/materials are required in addition to the life cycle analyses of the whole process. These interdisciplinary works and process systems engineering are important for the establishment of the DME economy, which is worthy of being considered a long-term research direction.

Author Contributions

Writing and editing—original draft preparation, C.L.; writing—review and supervision, Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work is financed by the National Natural Science Foundation of China (21603135), the Fundamental Research Funds for the Central Universities (GK201901001) and the Natural science basic research program of Shaanxi province (2021JQ-306).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, Y.; Deng, D.; Bao, X. Catalysis for Selected C1 Chemistry. Chem 2020, 6, 2497–2514. [Google Scholar] [CrossRef]
  2. Zhou, W.; Cheng, K.; Kang, J.; Zhou, C.; Subramanian, V.; Zhang, Q.; Wang, Y. New horizon in C1 chemistry: Breaking the selectivity limitation in transformation of syngas and hydrogenation of CO2 into hydrocarbon chemicals and fuels. Chem. Soc. Rev. 2019, 48, 3193–3228. [Google Scholar] [CrossRef] [PubMed]
  3. Hepburn, C.; Adlen, E.; Beddington, J.; Carter, E.A.; Fuss, S.; Dowell, N.M.; Minx, J.C.; Smith, P.; Williams, C.K. The technological and economic prospects for CO2 utilization and removal. Nature 2019, 575, 87–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Zhang, Z.; Wang, T.; Blunt, M.J.; Anthony, E.J.; Park, A.-H.A.; Hughes, R.W.; Webley, P.A.; Yan, J. Advances in carbon capture, utilization and storage. Appl. Energy 2020, 278, 115627. [Google Scholar] [CrossRef]
  5. Tapia, J.; Lee, J.; Ooi, R.; Foo, D.; Tan, R.R. A review of optimization and decision-making models for the planning of capture, utilization and storage (CCUS) systems. Sustain. Prod. Consum. 2018, 13, 1–15. [Google Scholar] [CrossRef]
  6. Chirone, R.; Paulillo, A.; Coppola, A.; Scala, F. Carbon capture and utilization via calcium looping, sorption enhanced methanation and green hydrogen: A techno-economic analysis and life cycle assessment study. Fuel 2022, 328, 125255. [Google Scholar] [CrossRef]
  7. Hernandez, E.; Hospital-Benito, D.; Moya, C.; Ortiz, R.; Belinchon, A.; Paramio, C.; Lemus, J.; Navarro, P.; Palomar, J. Integrated carbon capture and utilization based on bifunctional ionic liquids to save energy and emissions. Chem. Eng. J. 2022, 446, 137166. [Google Scholar] [CrossRef]
  8. Dimitriou, I.; Garcia-Gutierrez, P.; Elder, R.H.; Cuellar-Franca, R.M.; Azapagic, A.; Allena, R. Carbon dioxide utilisation for production of transport fuels: Process and economic analysis. Energy Environ. Sci. 2015, 8, 1775–1789. [Google Scholar] [CrossRef] [Green Version]
  9. Saeidi, S.; Amin, N.; Rahimpour, M.R. Hydrogenation of CO2 to value-added products-A review and potential future developments. J. CO2 Util. 2014, 5, 66–81. [Google Scholar] [CrossRef]
  10. Fan, W.K.; Tahir, M. Structured clay minerals-based nanomaterials for sustainable photo/thermal carbon dioxide conversion to cleaner fuels: A critical review. Sci. Total Environ. 2022, 845, 157206. [Google Scholar] [CrossRef]
  11. Goeppert, A.; Czaun, M.; Jones, J.; Prakash, G.K.; Olah, G.A. Recycling of carbon dioxide to methanol and derived products-closing the loop. Chem. Soc. Rev. 2014, 43, 7995–8048. [Google Scholar] [CrossRef] [PubMed]
  12. Olah, G.A.; Prakash, G.K.; Goeppert, A. Anthropogenic Chemical Carbon Cycle for a Sustainable Future. J. Am. Chem. Soc. 2011, 133, 12881–12898. [Google Scholar] [CrossRef] [PubMed]
  13. Olah, G.A. Beyond Oil and Gas: The Methanol Economy. Angew. Chem. Int. Ed. 2005, 44, 2636–2639. [Google Scholar] [CrossRef] [PubMed]
  14. Al-musleh, E.I.; Mallapragada, D.S.; Agrawal, R. Continuous power supply from a baseload renewable power plant. Appl. Energy 2014, 122, 83–93. [Google Scholar] [CrossRef]
  15. van Kampen, J.; Booneveld, S.; Boon, J.; Vente, J.; van Sint Annaland, M. Experimental validation of pressure swing regeneration for faster cycling in sorption enhanced dimethyl ether synthesis. Chem. Commun. 2020, 56, 13540–13541. [Google Scholar] [CrossRef]
  16. Gogate, M.R. The direct, one-step process for synthesis of dimethyl ether from syngas. III. DME as a chemical feedstock. Petrol. Sci. Technol. 2018, 36, 562–568. [Google Scholar] [CrossRef]
  17. Leonzio, G. State of art and perspectives about the production of methanol, dimethyl ether and syngas by carbon dioxide hydrogenation. J. CO2 Util. 2018, 27, 326–354. [Google Scholar] [CrossRef]
  18. Banivaheb, S.; Pitter, S.; Delgado, K.; Rubin, M.; Sauer, J.; Dittmeyer, R. Recent Progress in Direct DME Synthesis and Potential of Bifunctional Catalysts. Chem. Ing. Tech. 2022, 94, 240–255. [Google Scholar] [CrossRef]
  19. Ateka, A.; Rodriguez-Vega, P.; Erena, J.; Aguayo, A.; Bilbao, J. A review on the valorization of CO2: Focusing on the thermodynamics and catalyst design studies of the direct synthesis of dimethyl ether. Fuel Process. Technol. 2022, 233, 107310. [Google Scholar] [CrossRef]
  20. Barroso-Martin, I.; Infantes-Molina, A.; Jafarian-Fini, F.; Ballesteros-Plata, D.; Rodriguez-Castellon, E.; Moretti, E. Silica-Related Catalysts for CO2 Transformation into Methanol and Dimethyl Ether. Catalysts 2020, 10, 1282. [Google Scholar] [CrossRef]
  21. Vu, T.; Desgagnes, A.; Iliuta, M. Efficient approaches to overcome challenges in material development for conventional and intensified CO2 catalytic hydrogenation to CO, methanol, and DME. Appl. Catal. A Gen. 2021, 617, 118119. [Google Scholar] [CrossRef]
  22. Catizzone, E.; Freda, C.; Braccio, G.; Frusteri, F.; Bonura, G. Dimethyl ether as circular hydrogen carrier: Catalytic aspects of hydrogenation/dehydrogenation steps. J. Energy Chem. 2021, 58, 55–77. [Google Scholar] [CrossRef]
  23. Mota, N.; Ordonez, E.; Pawelec, B.; Fierro, J.; Navarro, R. Direct Synthesis of Dimethyl Ether from CO2: Recent Advances in Bifunctional/Hybrid Catalytic Systems. Catalysts 2021, 11, 411. [Google Scholar] [CrossRef]
  24. Roy, S.; Cherevotan, A.; Peter, S. Thermochemical CO2 Hydrogenation to Single Carbon Products: Scientific and Technological Challenges. ACS Energy Lett. 2018, 3, 1938–1966. [Google Scholar] [CrossRef] [Green Version]
  25. Puliyalil, H.; Jurkovic, D.; Dasireddy, V.; Likozar, B. A review of plasma-assisted catalytic conversion of gaseous carbon dioxide and methane into valueadded platform chemicals and fuels. RSC Adv. 2018, 8, 27481–27508. [Google Scholar] [CrossRef] [Green Version]
  26. Ateka, A.; Erena, J.; Bilbao, J.; Aguayo, A. Strategies for the Intensification of CO2 Valorization in the One-Step Dimethyl Ether Synthesis Process. Ind. Eng. Chem. Res. 2020, 59, 713–722. [Google Scholar] [CrossRef]
  27. Mondal, U.; Yadav, G. Perspective of dimethyl ether as fuel: Part I. Catalysis. J. CO2 Util. 2019, 32, 299–320. [Google Scholar] [CrossRef]
  28. Sobczak, J.; Wysocka, I.; Murgrabia, S.; Rogala, A. A Review on Deactivation and Regeneration of Catalysts for Dimethyl Ether Synthesis. Energies 2022, 15, 5420. [Google Scholar] [CrossRef]
  29. Olah, G.; Goeppert, A.; Prakash, G. Chemical Recycling of Carbon Dioxide to Methanol and Dimethyl Ether: From Greenhouse Gas to Renewable, Environmentally Carbon Neutral Fuels and Synthetic Hydrocarbons. J. Org. Chem. 2009, 74, 487–498. [Google Scholar] [CrossRef]
  30. Alvarez, A.; Bansode, A.; Urakawa, A.; Bavykina, A.; Wezendonk, T.; Makkee, M.; Gascon, J.; Kapteijn, F. Challenges in the Greener Production of Formates/Formic Acid, Methanol, and DME by Heterogeneously Catalyzed CO2 Hydrogenation Processes. Chem. Rev. 2017, 117, 9804–9838. [Google Scholar] [CrossRef]
  31. Pechenkin, A.; Potemkin, D.; Badmaev, S.; Smirnova, E.; Cherednichenko, K.; Vinokurov, V.; Glotov, A. CO2 hydrogenation to dimethyl ether over In2O3 catalysts supported on aluminosilicate halloysite nanotubes. Green Process. Syn. 2021, 10, 594–605. [Google Scholar] [CrossRef]
  32. Ni, Y.; Chen, Z.; Fu, Y.; Liu, Y.; Zhu, W.; Liu, Z. Selective conversion of CO2 and H2 into aromatics. Nat. Commun. 2019, 9, 1026–1032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Liu, C.; Kang, J.; Huang, Z.-Q.; Song, Y.-H.; Xiao, Y.-S.; Song, J.; He, J.-X.; Chang, C.-R.; Ge, H.-Q.; Wang, Y.; et al. Gallium nitride catalyzed the direct hydrogenation of carbon dioxide to dimethyl ether as primary product. Nat. Commun. 2021, 12, 2305. [Google Scholar] [CrossRef] [PubMed]
  34. Kurzina, I.A.; Reshetnikov, S.I.; Karakchieva, N.I.; Kurina, L.N. Direct synthesis of dimethyl ether from synthesis gas: Experimental study and mathematical modeling. Chem. Eng. J. 2017, 329, 135–141. [Google Scholar] [CrossRef]
  35. Shen, W.-J.; Jun, K.-W.; Choi, H.-S.; Lee, K.-W. Thermodynamic Investigation of Methanol and Dimethyl Ether Synthesis from CO2 Hydrogenation. Kor. J. Chem. Eng. 2000, 17, 210–216. [Google Scholar] [CrossRef]
  36. Jia, G.; Tan, Y.; Han, Y. A Comparative Study on the Thermodynamics of Dimethyl Ether Synthesis from CO Hydrogenation and CO2 Hydrogenation. Ind. Eng. Chem. Res. 2006, 45, 1152–1159. [Google Scholar] [CrossRef]
  37. Stangeland, K.; Li, H.; Yu, Z. Thermodynamic Analysis of Chemical and Phase Equilibria in CO2 Hydrogenation to Methanol, Dimethyl Ether, and Higher Alcohols. Ind. Eng. Chem. Res. 2018, 57, 4081–4094. [Google Scholar] [CrossRef]
  38. Swapnesh, A.; Srivastava, V.C.; Mall, I.D. Comparative Study on Thermodynamic Analysis of CO2 Utilization Reactions. Chem. Eng. Technol. 2014, 37, 1765–1777. [Google Scholar] [CrossRef]
  39. Ahmad, K.; Upadhyayula, S. Greenhouse Gas CO2 Hydrogenation to Fuels: A Thermodynamic Analysis. Environ. Prog. Sustain. Energy 2019, 38, 98–111. [Google Scholar] [CrossRef] [Green Version]
  40. Prachumsai, W.; Pangtaisong, S.; Assabumrungrat, S.; Bunruam, P.; Nakvachiratrakul, C.; Saebea, D.; Praserthdam, P.; Soisuwan, S. Carbon dioxide reduction to synthetic fuel on zirconia supported copper-based catalysts and gibbs free energy minimization: Methanol and dimethyl ether synthesis. J. Environ. Chem. Eng. 2021, 9, 104979. [Google Scholar] [CrossRef]
  41. Catizzone, E.; Bonura, G.; Migliori, M.; Frusteri, F.; Giordano, G. CO2 Recycling to Dimethyl Ether: State-of-the-Art and Perspectives. Molecules 2018, 23, 31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Palomo, J.; Rodriguez-Cano, M.A.; Rodríguez-Mirasol, J.; Cordero, T. On the kinetics of methanol dehydration to dimethyl ether on Zr-loaded P-containing mesoporous activated carbon catalyst. Chem. Eng. J. 2019, 378, 122198. [Google Scholar] [CrossRef]
  43. Azizia, Z.; Rezaeimanesh, M.; Tohidian, T.; Rahimpour, M.R. Dimethyl ether: A review of technologies and production challenges. Chem. Eng. Process. 2014, 82, 150–172. [Google Scholar] [CrossRef]
  44. Baltes, C.; Vukojevic, S.; Schüth, F. Correlations between synthesis, precursor, and catalyst structure and activity of a large set of CuO/ZnO/Al2O3 catalysts for methanol synthesis. J. Catal. 2008, 258, 334–344. [Google Scholar] [CrossRef]
  45. Robbins, J.L.; Iglesia, E.; Kelkar, C.P.; DeRites, B. Methanol synthesis over Cu/SiO2 catalysts. Catal. Lett. 1991, 10, 1–10. [Google Scholar] [CrossRef]
  46. Dong, X.; Li, F.; Zhao, N.; Xiao, F.; Wang, J.; Tan, Y. CO2 hydrogenation to methanol over Cu/ZnO/ZrO2 catalysts prepared by precipitation-reduction method. Appl. Catal. B Environ. 2016, 191, 8–17. [Google Scholar] [CrossRef]
  47. Natesakhawat, S.; Lekse, J.W.; Baltrus, J.P.; Ohodnicki, J.P.R.; Howard, B.H.; Deng, X.; Matranga, C. Active Sites and Structure−Activity Relationships of Copper-Based Catalysts for Carbon Dioxide Hydrogenation to Methanol. ACS Catal. 2012, 2, 1667–1676. [Google Scholar] [CrossRef]
  48. Zhou, G.; He, Z.; Dong, X. Role of Metal Oxides in Cu-Based Catalysts with NaBH4 Reduction for the Synthesis of Methanol from CO2/H2. Catal. Lett. 2021, 151, 1091–1101. [Google Scholar] [CrossRef]
  49. Zhang, F.; Liu, Y.; Xu, X.; Yang, P.; Miao, P.; Zhang, Y.; Sun, Q. Effect of Al-containing precursors on Cu/ZnO/Al2O3 catalyst for methanol production. Fuel Process. Technol. 2018, 178, 148–155. [Google Scholar] [CrossRef]
  50. Gao, P.; Xie, R.; Wang, H.; Zhong, L.; Xia, L.; Zhang, Z.; Wei, W.; Sun, Y. Cu/Zn/Al/Zr catalysts via phase-pure hydrotalcite-like compounds for methanol synthesis from carbon dioxide. J. CO2 Util. 2015, 11, 41–48. [Google Scholar] [CrossRef]
  51. Behrens, M.; Zander, S.; Kurr, P.; Jacobsen, N.; Senker, J.; Koch, G.; Ressler, T.; Fischer, R.W.; Schlogl, R. Performance Improvement of Nanocatalysts by Promoter-Induced Defects in the Support Material: Methanol Synthesis over Cu/ZnO:Al. J. Am. Chem. Soc. 2013, 135, 6061–6068. [Google Scholar] [CrossRef] [PubMed]
  52. Ren, H.; Xu, C.-H.; Zhao, H.-Y.; Wang, Y.-X.; Liu, J.; Liu, J.-Y. Methanol synthesis from CO2 hydrogenation over Cu/γ-Al2O3 catalysts modified by ZnO, ZrO2 and MgO. J. Ind. Eng. Chem. 2015, 28, 261–267. [Google Scholar] [CrossRef]
  53. Din, I.; Shaharun, M.; Subbarao, D.; Naeem, A. Synthesis, characterization and activity pattern of carbon nanofibers based copper/zirconia catalysts for carbon dioxide hydrogenation to methanol: Influence of calcination temperature. J. Power Sources 2015, 274, 619–628. [Google Scholar] [CrossRef]
  54. Lei, H.; Nie, R.; Wu, G.; Hou, Z. Hydrogenation of CO2 to CH3OH over Cu/ZnO catalysts with different ZnO morphology. Fuel 2015, 154, 161–166. [Google Scholar] [CrossRef]
  55. Gao, P.; Zhong, L.; Zhang, L.; Wang, H.; Zhao, N.; Wei, W.; Sun, Y. Yttrium oxide modified Cu/ZnO/Al2O3 catalysts via hydrotalcite-like precursors for CO2 Hydrogenation to methanol. Catal. Sci. Technol. 2015, 5, 4365–4377. [Google Scholar] [CrossRef]
  56. Ahouari, H.; Soualah, A.; Le Valant, A.; Pinard, L.; Magnoux, P.; Pouilloux, Y. Methanol synthesis from CO2 hydrogenation over copper based catalysts. React. Kinet. Mech. Catal. 2013, 110, 131–145. [Google Scholar] [CrossRef]
  57. Li, K.; Chen, J.G. CO2 Hydrogenation to Methanol over ZrO2-Containing Catalysts: Insights into ZrO2 Induced Synergy. ACS Catal. 2019, 9, 7840–7861. [Google Scholar] [CrossRef]
  58. Wang, Y.; Kattel, S.; Gao, W.; Li, K.; Liu, P.; Chen, J.G.; Wang, H. Exploring the ternary interactions in Cu-ZnO-ZrO2 catalysts for efficient CO2 hydrogenation to methanol. Nat. Commun. 2019, 10, 1166. [Google Scholar] [CrossRef] [Green Version]
  59. Stangeland, K.; Navarro, H.H.; Huynh, H.L.; Tucho, W.M.; Yu, Z. Tuning the interfacial sites between copper and metal oxides (Zn, Zr, In) for CO2 hydrogenation to methanol. Chem. Eng. Sci. 2021, 238, 116603. [Google Scholar] [CrossRef]
  60. Frei, E.; Gaur, A.; Lichtenberg, H.; Heine, C.; Friedrich, M.; Greiner, M.; Lunkenbein, T.; Grunwaldt, J.; Schlögl, R. Activating a Cu/ZnO:Al Catalyst-much more than Reduction: Decomposition, Self-doping and Polymorphism. Chemcatchem 2019, 11, 1587–1592. [Google Scholar] [CrossRef]
  61. Gao, P.; Li, F.; Xiao, F.; Zhao, N.; Wei, W.; Zhong, L.; Sun, Y. Effect of hydrotalcite-containing precursors on the performance of Cu/Zn/Al/Zr catalysts for CO2 hydrogenation: Introduction of Cu2+ at different formation stages of precursors. Catal. Today 2012, 194, 9–15. [Google Scholar] [CrossRef]
  62. Dasireddy, V.; Likozar, B. The role of copper oxidation state in Cu/ZnO/Al2O3 catalysts in CO2 hydrogenation and methanol productivity. Renew. Energy 2019, 140, 452–460. [Google Scholar] [CrossRef]
  63. Stangeland, K.; Chamssine, F.; Fu, W.; Huang, Z.; Duan, X.; Yu, Z. CO2 hydrogenation to methanol over partially embedded Cu within Zn-Al oxide and the effect of indium. J. CO2 Util. 2021, 50, 101609. [Google Scholar] [CrossRef]
  64. Zhang, Z.; You, R.; Huang, W. Cu2O Nanocrystal Model Catalysts. Chin. J. Chem. 2022, 40, 846–855. [Google Scholar] [CrossRef]
  65. Pu, Y.; Luo, Y.; Wei, X.; Sun, J.; Li, L.; Zou, W.; Dong, L. Synergistic effects of Cu2O-decorated CeO2 on photocatalytic CO2 reduction: Surface Lewis acid/base and oxygen defect. Appl. Catal. B Environ. 2019, 254, 580–586. [Google Scholar] [CrossRef]
  66. Zou, T.; Araújo, T.P.; Krumeich, F.; Mondelli, C.; Pérez-Ramírez, J. ZnO-Promoted Inverse ZrO2-Cu Catalysts for CO2-Based Methanol Synthesis under Mild Conditions. ACS Sustain. Chem. Eng. 2022, 10, 81–90. [Google Scholar] [CrossRef]
  67. Pasupulety, N.; Al-Zahrani, A.A.; Daous, M.A.; Podila, S.; Driss, H. A study on highly active Cu-Zn-Al-K catalyst for CO2 hydrogenation to methanol. Arab. J. Chem. 2021, 14, 102951. [Google Scholar] [CrossRef]
  68. Chen, S.; Zhang, J.; Song, F.; Zhang, Q.; Yang, G.; Zhang, M.; Wang, X.; Xie, H.; Tan, Y. Induced high selectivity methanol formation during CO2 hydrogenation over a CuBr2-modified CuZnZr catalyst. J. Catal. 2020, 389, 47–59. [Google Scholar] [CrossRef]
  69. Nakamura, J.; Choia, Y.; Fujitan, T. On the issue of the active site and the role of ZnO in Cu/ZnO methanol synthesis catalysts. Top Catal. 2003, 22, 277–284. [Google Scholar] [CrossRef]
  70. Studt, F.; Behrens, M.; Kunkes, E.L.; Thomas, N.; Zander, S.; Tarasov, A.; Schumann, J.; Frei, E.; Varley, J.B.; Abild-Pedersen, F.; et al. The Mechanism of CO and CO2 Hydrogenation to Methanol over Cu-Based Catalysts. Chemcatchem 2015, 7, 1105–1111. [Google Scholar] [CrossRef]
  71. Kattel, S.; Ramirez., P.J.; Chen, J.; Rodriguez, J.; Liu, P. Active sites for CO2 hydrogenation to methanol on Cu/ZnO catalysts. Science 2017, 355, 1296–1299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Behrens, M.; Studt, F.; Kasatkin, I.; Kühl, S.; Hävecker, M.; Abild-Pedersen, F.; Zander, S.; Girgsdies, F.; Kurr, P.; Kniep, B.-L.; et al. The Active Site of Methanol Synthesis over Cu/ZnO/Al2O3 Industrial Catalysts. Science 2012, 336, 893–897. [Google Scholar] [CrossRef] [PubMed]
  73. Kuld, S.; Conradsen, C.; Moses, P.G.; Chorkendorff, I.; Sehested, J. Quantification of Zinc Atoms in a Surface Alloy on Copper in an Industrial-Type Methanol Synthesis Catalyst. Angew. Chem. Int. Ed. 2014, 53, 5941–5945. [Google Scholar] [CrossRef] [PubMed]
  74. Du, J.; Zhang, Y.; Wang, K.; Ding, F.; Jia, S.; Liu, G.; Tan, L. Investigation on the promotional role of Ga2O3 on the CuO-ZnO/HZSM-5 catalyst for CO2 hydrogenation. RSC Adv. 2021, 11, 14426. [Google Scholar] [CrossRef]
  75. Chen, K.; Fang, H.; Wu, S.; Liu, X.; Zheng, J.; Zhou, S.; Duan, X.; Zhuang, Y.; Tsang, S.C.E.; Yuan, Y. CO2 hydrogenation to methanol over Cu catalysts supported on La-modified SBA-15: The crucial role of Cu-LaOx interfaces. Appl. Catal. B Environ. 2019, 251, 119–129. [Google Scholar] [CrossRef]
  76. Liu, C.; Liu, P. Mechanistic Study of Methanol Synthesis from CO2 and H2 on a Modified Model Mo6S8 Cluster. ACS Catal. 2015, 5, 1004–1012. [Google Scholar] [CrossRef]
  77. Wang, W.; Zhang, X.; Guo, M.; Li, J.; Peng, C. An investigation of the CH3OH and CO selectivity of CO2 hydrogenation over Cu-Ce-Zr catalysts. Front. Chem. Sci. Eng. 2022, 16, 950–962. [Google Scholar] [CrossRef]
  78. Tang, C.; Tang, S.; Sha, F.; Han, Z.; Feng, Z.; Wang, J.; Li, C. Insights into the Selectivity Determinant and Rate-Determining Step of CO2 Hydrogenation to Methanol. J. Phys. Chem. C 2022, 126, 10399–10407. [Google Scholar] [CrossRef]
  79. Wang, X.; Zhang, H. Kinetically relevant variation triggered by hydrogen pressure: A mechanistic case study of CO2 hydrogenation to methanol over Cu/ZnO. J. Catal. 2022, 406, 145–156. [Google Scholar] [CrossRef]
  80. Chinchen, G.C.; Denny, P.J.; Parke, D.G. Mechanism of methanol synthesis from CO2/CO/H2 mixtures over copper/zinc oxide/alumina catalysts: Use of 14C-labelled reactants. Appl. Catal. 1987, 30, 333–338. [Google Scholar] [CrossRef]
  81. Larmier, K.; Liao, W.C.; Tada, S.; Lam, E.; Verel, R.; Bansode, A.; Urakawa, A.; Comas-Vives, A.; Coperet, C. CO2-to-Methanol Hydrogenation on Zirconia-Supported Copper Nanoparticles: Reaction Intermediates and the Role of the Metal-Support Interface. Angew. Chem. Int. Ed. 2017, 56, 2318–2323. [Google Scholar] [CrossRef] [PubMed]
  82. Kattel, S.; Yan, B.; Yang, Y.; Chen, J.G.; Liu, P. Optimizing Binding Energies of Key Intermediates for CO2 Hydrogenation to Methanol over Oxide-Supported Copper. J. Am. Chem. Soc. 2016, 138, 12440–12450. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, L.; Fan, F.; Bai, M.; Xue, F.; Ma, X.; Jiang, Z.; Fang, T. Mechanistic study of methanol synthesis from CO2 hydrogenation on Rhdoped Cu(111) surfaces. Mol. Catal. 2019, 466, 26–36. [Google Scholar] [CrossRef]
  84. Liu, L.; Wang, X.; Lu, S.; Li, J.; Zhang, H.; Su, X.; Xue, F.; Cao, B.; Fang, T. Mechanism of CO2 hydrogenation over a Zr1-Cu single-atom catalyst. New J. Chem. 2022, 46, 5043–5051. [Google Scholar] [CrossRef]
  85. Yu, J.; Yang, M.; Zhang, J.; Ge, Q.; Zimina, A.; Pruessmann, T.; Zheng, L.; Grunwaldt, J.-D.; Sun, J. Stabilizing Cu+ in Cu/SiO2 Catalysts with a Shattuckite-Like Structure Boosts CO2 Hydrogenation into Methanol. ACS Catal. 2020, 10, 14694–14706. [Google Scholar] [CrossRef]
  86. Kunkes, E.; Studt, F.; Abild-Pedersen, F.; Schlögl, R.; Behrens, M. Hydrogenation of CO2 to methanol and CO on Cu/ZnO/Al2O3: Is there a common intermediate or not? J. Catal. 2015, 328, 43–48. [Google Scholar] [CrossRef] [Green Version]
  87. Poto, S.; van Berkel, D.; Gallucci, F.; Neira d’Angelo, M. Kinetic modelling of the methanol synthesis from CO2 and H2 over a CuO/CeO2/ZrO2 catalyst: The role of CO2 and CO hydrogenation. Chem. Eng. J. 2022, 435, 134946. [Google Scholar] [CrossRef]
  88. Ren, S.; Fan, X.; Shang, Z.; Shoemaker, W.R.; Ma, L.; Wu, T.; Li, S.; Klinghoffer, N.B.; Yu, M.; Liang, X. Enhanced Catalytic Performance of Zr-Modified CuO/ZnO/Al2O3 Catalyst for Methanol and DME Synthesis via CO2 Hydrogenation. J. CO2 Util. 2020, 36, 82–95. [Google Scholar] [CrossRef]
  89. Pedram, M.Z.; Kazemeini, M.; Fattahi, M.; Amjadian, A. A Physicochemical Evaluation of Modified HZSM-5 Catalyst Utilized for Production of Dimethyl Ether from Methanol. Pet. Sci. Technol. 2014, 32, 904–911. [Google Scholar] [CrossRef]
  90. Malik, A.; Zaman, S.; Al-Zahrani, A.; Daous, M. Turning CO2 into di-methyl ether (DME) using Pd based catalysts-Role of Ca in tuning the activity and selectivity. J. Ind. Eng. Chem. 2021, 103, 67–79. [Google Scholar] [CrossRef]
  91. Oyola-Rivera, O.; Baltanas, M.; Cardona-Martinez, N. CO2 hydrogenation to methanol and dimethyl ether by Pd-Pd2Ga catalysts supported over Ga2O3 polymorphs. J. CO2 Util. 2015, 9, 8–15. [Google Scholar] [CrossRef]
  92. Liao, F.; Benedict Lo, T.; Tsang, S. Recent Developments in Palladium-Based Bimetallic Catalysts. ChemCatChem 2015, 7, 1998–2014. [Google Scholar] [CrossRef]
  93. Chu, R.; Song, C.; Hou, W.; Meng, X.; Miao, Z.; Li, X.; Wu, G.; Wan, Y.; Bai, L. Improved stability of Pd/HZSM-5 bifunctional catalysts by the addition of promoters (CeO2, CaO) for the One-step synthesis of dimethyl ether from Sulfur-containing CO2 hydrogenation. J. Taiwan Instit. Chem. Eng. 2017, 80, 1041–1047. [Google Scholar] [CrossRef]
  94. Collins, S.E.; Baltanas, M.A.; Bonivardi, A.L. An infrared study of the intermediates of methanol synthesis from carbon dioxide over Pd/β-Ga2O3. J. Catal. 2004, 226, 410–421. [Google Scholar] [CrossRef]
  95. Schild, C.; Wokaun, A.; Baiker, A. On the mechanism of CO and CO2 hydrogenation reactions on zirconia-supported catalysts: A diffuse reflectance FTIR study: Part I. Identification of surface species and methanation reactions on palladium/zirconia catalysts. J. Mol. Catal. 1990, 63, 223–242. [Google Scholar] [CrossRef]
  96. Arunajatesan, V.; Subramaniam, B.; Hutchenson, K.W.; Herkes, F. In-situ FTIR investigations of reverse water gas shift reaction activity at supercritical conditions. Chem. Eng. Sci. 2007, 62, 5062–5069. [Google Scholar] [CrossRef]
  97. Malik, A.S.; Zaman, S.F.; Al-Zahrani, A.A.; Daous, M.A.; Driss, H.; Petrov, L.A. Development of highly selective PdZn/CeO2 and Ca-doped PdZn/CeO2 catalysts for methanol synthesis from CO2 hydrogenation. Appl. Catal. A Gen. 2018, 560, 42–53. [Google Scholar] [CrossRef]
  98. Cabilla, G.C.; Bonivardi, A.L.; Baltanas, M.A. FTIR Study of the Adsorption of Methanol on Clean and Ca-Promoted Pd/SiO2 Catalysts. J. Catal. 2001, 201, 213–220. [Google Scholar] [CrossRef]
  99. Gao, W.; Wang, H.; Wang, Y.; Guo, W.; Jia, M. Dimethyl ether synthesis from CO2 hydrogenation on La-modified CuO-ZnO-Al2O3/HZSM-5 bifunctional catalysts. J. Rare Earth. 2013, 31, 470–475. [Google Scholar] [CrossRef]
  100. Frusteri, F.; Migliori, M.; Cannilla, C.; Frusteri, L.; Catizzone, E.; Aloiseb, A.; Giordano, G.; Bonura, G. Direct CO2-to-DME hydrogenation reaction: New evidences of a superior behaviour of FER-based hybrid systems to obtain high DME yield. J. CO2 Util. 2017, 18, 353–361. [Google Scholar] [CrossRef]
  101. Frusteri, F.; Cordaro, M.; Cannilla, C.; Bonura, G. Multifunctionality of Cu-ZnO-ZrO2/H-ZSM5 catalysts for the one-step CO2-to-DME hydrogenation reaction. Appl. Catal. B Environ. 2015, 162, 57–65. [Google Scholar] [CrossRef]
  102. Bonura, G.; Cannilla, C.; Frusteri, L.; Frusteri, F. The influence of different promoter oxides on the functionality of hybrid CuZn-ferrierite systems for the production of DME from CO2-H2 mixtures. Appl. Catal. A Gen. 2017, 544, 21–29. [Google Scholar] [CrossRef]
  103. Cara, C.; Secci, F.; Lai, S.; Mameli, V.; Skrodczky, K.; Russo, P.A.; Ferrara, F.; Rombi, E.; Pinna, N.; Mureddu, M.; et al. On the design of mesostructured acidic catalysts for the one-pot dimethyl ether production from CO2. J. CO2 Util. 2022, 62, 102066. [Google Scholar] [CrossRef]
  104. Zhang, M.; Liu, Z.; Lin, G.; Zhang, H. Pd/CNT-promoted Cu-ZrO2/HZSM-5 hybrid catalysts for direct synthesis of DME from CO2/H2. Appl. Catal. A Gen. 2013, 451, 28–35. [Google Scholar] [CrossRef]
  105. Navarro-Jaen, S.; Virginie, M.; Morin, J.; Thuriot-Roukos, J.; Wojcieszak, R.; Khodakov, A. Hybrid monometallic and bimetallic copper–palladium zeolite catalysts for direct synthesis of dimethyl ether from CO2. New J. Chem. 2022, 46, 3889–3900. [Google Scholar] [CrossRef]
  106. Krim, K.; Sachse, A.; Le Valant, A.; Pouilloux, Y.; Hocine, S. One Step Dimethyl Ether (DME) Synthesis from CO2 Hydrogenation over Hybrid Catalysts Containing Cu/ZnO/Al2O3 and Nano-Sized Hollow ZSM-5 Zeolites. Catal. Lett. 2022, 1–12. [Google Scholar] [CrossRef]
  107. Hu, Y.; Zhang, Y.; Du, J.; Li, C.; Wang, K.; Liu, L.; Yu, X.; Wang, K.; Liu, N. The influence of composition on the functionality of hybrid CuO-ZnO-Al2O3/HZSM-5 for the synthesis of DME from CO2 hydrogenation. RSC Adv. 2018, 8, 30387–30395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Liu, R.; Tian, H.; Yang, A.; Zha, F.; Ding, J.; Chang, Y. Preparation of HZSM-5 membrane packed CuO-ZnO-Al2O3 nanoparticles for catalysing carbon dioxide hydrogenation to dimethyl ether. Appl. Surf. Sci. 2015, 345, 1–9. [Google Scholar] [CrossRef]
  109. Martin, O.; Martin, A.J.; Mondelli, C.; Mitchell, S.; Segawa, T.F.; Hauert, R.; Drouilly, C.; Curulla-Ferre, D.; Perez-Ramirez, J. Indium Oxide as a Superior Catalyst for Methanol Synthesis by CO2 Hydrogenation. Angew. Chem. Int. Ed. 2016, 55, 6261–6265. [Google Scholar] [CrossRef]
  110. Gao, P.; Li, S.; Bu, X.; Dang, S.; Liu, Z.; Wang, H.; Zhong, L.; Qiu, M.; Yang, C.; Cai, J.; et al. Direct conversion of CO2 into liquid fuels with high selectivity over a bifunctional catalyst. Nat. Chem. 2017, 9, 1019–1024. [Google Scholar] [CrossRef]
  111. Ye, J.; Liu, C.; Mei, D.; Ge, Q. Active Oxygen Vacancy Site for Methanol Synthesis from CO2 Hydrogenation on In2O3: A DFT Study. ACS Catal. 2013, 3, 1296–1306. [Google Scholar] [CrossRef]
  112. Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjær, C.F.; Hummelshøj, J.S.; Dahl, S.; Chorkendorff, I.; Nørskov, J.K. Discovery of a Ni-Ga catalyst for carbon dioxide reduction to methanol. Nat. Chem. 2014, 6, 320–324. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, X.; Wang, M.; Yin, H.; Hu, J.; Cheng, K.; Kang, J.; Zhang, Q.; Wang, Y. Tandem Catalysis for Hydrogenation of CO and CO2 to Lower Olefins with Bifunctional Catalysts Composed of Spinel Oxide and SAPO-34. ACS Catal. 2020, 10, 8303–8314. [Google Scholar] [CrossRef]
  114. Zhang, X.; Zhang, G.; Liu, W.; Yuan, F.; Wang, J.; Zhu, J.; Jiang, X.; Zhang, A.; Ding, F.; Song, C.; et al. Reaction-driven surface reconstruction of ZnAl2O4 boosts the methanol selectivity in CO2 catalytic hydrogenation. Appl. Catal. B Environ. 2021, 284, 119700. [Google Scholar] [CrossRef]
  115. Saravanan, K.; Ham, H.; Tsubaki, N.; Bae, J.W. Recent progress for direct synthesis of dimethyl ether from syngas on the heterogeneous bifunctional hybrid catalysts. Appl. Catal. B Environ. 2017, 217, 494–522. [Google Scholar] [CrossRef]
  116. Sierra, I.; Erena, J.; Aguayo, A.T.; Arandes, J.M.; Olazar, M.; Bilbao, J. Co-feeding water to attenuate deactivation of the catalyst metallic function (CuO-ZnO-Al2O3) by coke in the direct synthesis of dimethyl ether. Appl. Catal. B Environ. 2011, 106, 167–173. [Google Scholar] [CrossRef]
  117. Abu-Dahrieh, J.; Rooney, D.; Goguet, A.; Saih, Y. Activity and deactivation studies for direct dimethyl ether synthesis using CuO-ZnO-Al2O3 with NH4ZSM-5, HZSM-5 or γ-Al2O3. Chem. Eng. J. 2012, 203, 201–211. [Google Scholar] [CrossRef] [Green Version]
  118. Xu, M.; Lunsford, J.H.; Goodman, D.W.; Bhattacharyya, A. Synthesis of dimethyl ether (DME) from methanol over solid-acid catalysts. Appl. Catal. A Gen. 1997, 149, 289–301. [Google Scholar] [CrossRef]
  119. Sanchez-Contador, M.; Ateka, A.; Ibanez, M.; Bilbao, J.; Aguayo, A. Influence of the operating conditions on the behavior and deactivation of a CuO-ZnO-ZrO2@SAPO-11 core-shell-like catalyst in the direct synthesis of DME. Renew. Energy 2019, 138, 585–597. [Google Scholar] [CrossRef]
  120. Montesano, R.; Narvaez, A.; Chadwick, D. Shape-selectivity effects in syngas-to-dimethyl ether conversion over Cu/ZnO/Al2O3 and zeolite mixtures: Carbon deposition and by-product formation. Appl. Catal. A Gen. 2014, 482, 69–77. [Google Scholar] [CrossRef]
  121. Frusteri, F.; Bonura, G.; Cannilla, C.; Ferrante, G.; Aloise, A.; Catizzone, E.; Migliorib, M.; Giordano, G. Stepwise tuning of metal-oxide and acid sites of CuZnZr-MFI hybrid catalysts for the direct DME synthesis by CO2 hydrogenation. Appl. Catal. B Environ. 2015, 176–177, 522–531. [Google Scholar] [CrossRef]
  122. Zhou, X.; Su, T.; Jiang, Y.; Qin, Z.; Ji, H.; Guo, Z. CuO-Fe2O3-CeO2/HZSM-5 bifunctional catalyst hydrogenated CO2 for enhanced dimethyl ether synthesis. Chem. Eng. Sci. 2016, 153, 10–20. [Google Scholar] [CrossRef]
  123. Jin, D.; Zhu, B.; Hou, Z.; Fei, J.; Lou, H.; Zheng, X. Dimethyl ether synthesis via methanol and syngas over rare earth metals modified zeolite Y and dual Cu-Mn-Zn catalysts. Fuel 2007, 86, 2707–2713. [Google Scholar] [CrossRef]
  124. Aguayo, A.T.; Erena, J.; Sierra, I.; Olazar, M.; Bilbao, J. Deactivation and regeneration of hybrid catalysts in the single-step synthesis of dimethyl ether from syngas and CO2. Catal. Today 2005, 106, 265–270. [Google Scholar] [CrossRef]
  125. Hassanpour, S.; Taghizadeh, M.; Yaripour, F. Preparation, Characterization, and Activity Evaluation of H-ZSM-5 Catalysts in Vapor-Phase Methanol Dehydration to Dimethyl Ether. Ind. Eng. Chem. Res. 2010, 49, 4063–4069. [Google Scholar] [CrossRef]
  126. Catizzone, E.; Aloise, A.; Migliori, M.; Giordano, G. The effect of FER zeolite acid sites in methanol-to-dimethyl-ether catalytic dehydration. J. Energ. Chem. 2017, 26, 406–415. [Google Scholar] [CrossRef] [Green Version]
  127. Liu, X.; Zhou, W.; Yang, Y.; Cheng, K.; Kang, J.; Zhang, L.; Zhang, G.; Min, X.; Zhang, Q.; Wang, Y. Design of efficient bifunctional catalysts for direct conversion of syngas into lower olefins via methanol/dimethyl ether intermediates. Chem. Sci. 2018, 9, 4708–4718. [Google Scholar] [CrossRef] [Green Version]
  128. Catizzone, E.; Aloise, A.; Migliori, M.; Giordano, G. From 1-D to 3-D zeolite structures: Performance assessment in catalysis of vapour-phase methanol dehydration to DME. Micropor. Mesopor. Mat. 2017, 243, 102–111. [Google Scholar] [CrossRef]
  129. Kim, S.D.; Baek, S.C.; Lee, Y.; Jun, K.; Kim, M.J.; Yoo, I.S. Effect of γ-alumina content on catalytic performance of modified ZSM-5 for dehydration of crude methanol to dimethyl ether. Appl. Catal. A Gen. 2006, 309, 139–143. [Google Scholar] [CrossRef]
Figure 1. Scheme for the formation and utilization of methanol and dimethyl ether.
Figure 1. Scheme for the formation and utilization of methanol and dimethyl ether.
Catalysts 12 01375 g001
Figure 2. Equilibrium product yield in the CO2 hydrogenation. Reproduced with permission from Shen et al. [35], published by Springer Nature, 2000.
Figure 2. Equilibrium product yield in the CO2 hydrogenation. Reproduced with permission from Shen et al. [35], published by Springer Nature, 2000.
Catalysts 12 01375 g002
Figure 3. Thermodynamic results in the equilibrium reactions. (a) Equilibrium methanol conversion and DME selectivity in the methanol dehydration. (b) Equilibrium DME yield in the one-step or two-step CO2 hydrogenation. Reproduced with permission from Shen et al. [35] and Palomo et al. [42], respectively published by Springer Nature, 2000, and Elsevier, 2019.
Figure 3. Thermodynamic results in the equilibrium reactions. (a) Equilibrium methanol conversion and DME selectivity in the methanol dehydration. (b) Equilibrium DME yield in the one-step or two-step CO2 hydrogenation. Reproduced with permission from Shen et al. [35] and Palomo et al. [42], respectively published by Springer Nature, 2000, and Elsevier, 2019.
Catalysts 12 01375 g003
Figure 5. Energy (E) and standard Gibbs free-energy (G) pathway calculated for the CO2 hydrogenation to methanol over Cu/ZrO2. Reproduced with permission from Larmier et al. [81], published by John Wiley and Sons, 2017.
Figure 5. Energy (E) and standard Gibbs free-energy (G) pathway calculated for the CO2 hydrogenation to methanol over Cu/ZrO2. Reproduced with permission from Larmier et al. [81], published by John Wiley and Sons, 2017.
Catalysts 12 01375 g005
Figure 6. Schematic reaction mechanism of CO2 hydrogenation over Pd-based catalysts with addition of different oxides.
Figure 6. Schematic reaction mechanism of CO2 hydrogenation over Pd-based catalysts with addition of different oxides.
Catalysts 12 01375 g006
Figure 7. CO2 conversion (XCO2) and selectivity of dimethyl ether (SDME) with (square) and without (circle) H2S in the gas feed. (a) CaO–CeO2–Pd/HZSM-5; (b) Cu–ZnO–Al2O3/HZSM-5. Conditions: 250 °C, 3 MPa, 1600 mL·g−1·h−1, H2/CO2 = 4.4. Reproduced with permission from Chu et al. [93], published by Elsevier, 2017.
Figure 7. CO2 conversion (XCO2) and selectivity of dimethyl ether (SDME) with (square) and without (circle) H2S in the gas feed. (a) CaO–CeO2–Pd/HZSM-5; (b) Cu–ZnO–Al2O3/HZSM-5. Conditions: 250 °C, 3 MPa, 1600 mL·g−1·h−1, H2/CO2 = 4.4. Reproduced with permission from Chu et al. [93], published by Elsevier, 2017.
Catalysts 12 01375 g007
Figure 8. Catalytic performance over the In2O3/halloysite nanotubes catalysts at T = 200–300 °C, P = 4 MPa and GHSV = 12,000 mL·g−1·h−1. Reproduced with permission from Pechenkin et al. [31], published by De Gruyter, 2021.
Figure 8. Catalytic performance over the In2O3/halloysite nanotubes catalysts at T = 200–300 °C, P = 4 MPa and GHSV = 12,000 mL·g−1·h−1. Reproduced with permission from Pechenkin et al. [31], published by De Gruyter, 2021.
Catalysts 12 01375 g008
Figure 9. The density functional theory calculation results of CO2 hydrogenation over GaN. (a) Gibbs free energy diagram of the CO2 hydrogenation to methyl (CH3 *) on the GaN(100) plane. (b) Gibbs free energy diagram of the CO2 hydrogenation to formate (HCOO*) on the GaN(110) plane. (c) Gibbs free energy diagram for the coupling of HCOO* and CH3 * to CH3OCH3 * on the (110)/(100) interface. TS: transition states. Reproduced with permission from Liu et al. [33], published by Springer Nature, 2021.
Figure 9. The density functional theory calculation results of CO2 hydrogenation over GaN. (a) Gibbs free energy diagram of the CO2 hydrogenation to methyl (CH3 *) on the GaN(100) plane. (b) Gibbs free energy diagram of the CO2 hydrogenation to formate (HCOO*) on the GaN(110) plane. (c) Gibbs free energy diagram for the coupling of HCOO* and CH3 * to CH3OCH3 * on the (110)/(100) interface. TS: transition states. Reproduced with permission from Liu et al. [33], published by Springer Nature, 2021.
Catalysts 12 01375 g009
Figure 10. Catalytic performance over the ZnAlOx catalyst at T = 320 °C, P = 3 MPa and GHSV = 12,000 mL·g−1·h−1. Reproduced with permission from Ni et al. [32], published by Springer Nature, 2019.
Figure 10. Catalytic performance over the ZnAlOx catalyst at T = 320 °C, P = 3 MPa and GHSV = 12,000 mL·g−1·h−1. Reproduced with permission from Ni et al. [32], published by Springer Nature, 2019.
Catalysts 12 01375 g010
Figure 11. Scheme for the reaction pathway of the one-step CO2 hydrogenation to DME.
Figure 11. Scheme for the reaction pathway of the one-step CO2 hydrogenation to DME.
Catalysts 12 01375 g011
Table 1. Thermodynamic results of reactions involved in the CO2 hydrogenation to DME. Results are originated from [18,21,34].
Table 1. Thermodynamic results of reactions involved in the CO2 hydrogenation to DME. Results are originated from [18,21,34].
ReactionΔH298K (kJ·mol−1)ΔG298K (kJ·mol−1)
(CO2-to-methanol)
CO2 + 3H2 = CH3OH + H2O
−49.4+3.5
(CO2-to-DME)
2CO2 + 6H2 = CH3OCH3 + 3H2O
−122.0−5.0
(RWGS)
CO2 + H2 = CO + H2O
+41.2−28.6
(CO-to-methanol)
CO + 2H2 = CH3OH
−90.6−25.2
(CO-to-DME)
2CO + 4H2 = CH3OCH3 + H2O
−102.4−33.5
Table 2. Catalytic performance of Cu- and Pd-based catalysts in the CO2 hydrogenation to dimethyl ether (DME).
Table 2. Catalytic performance of Cu- and Pd-based catalysts in the CO2 hydrogenation to dimethyl ether (DME).
CatalystT (°C)P (MPa)GHSV (mL·g−1·h−1)CO2 Conversion (%)DME Selectivity 1 (%)STYDME 2 (mmol·g−1·h−1)Reference
Cu–ZnO–Al2O3/
HZSM-5
2253.048,0002.975.65.0[106]
Cu–ZnO–Al2O3/
HZSM-5
2603.0150019.291.61.7[107]
Cu–ZnO–Al2O3/
HZSM-5
3003.0180029.524.10.92[108]
Cu–ZnO–Al2O3/
Ferrierite
2503.012,00019.879.06.2[103]
Cu–ZnO–CeO2/
Ferrierite
2603.0880014.071.65.0[102]
Cu–ZnO–ZrO2/
Ferrierite
2605.0880026.081.37.2[100]
Cu–ZnO–ZrO2/HZSM-52403.010,00013.480.84.9[101]
La–Cu–ZnO–Al2O3/
HZSM-5
2503.0300043.894.39.1[99]
Cu–ZnO–Ga2O3/
HZSM-5
2603.0150022.389.42.8[74]
CaO–Pd/CeO23252.0240030.970.55.3[90]
Pd/Ga2O3/Nb2O52701.727005.053.00.39[91]
CeO2–CaO–Pd/HZSM-52503.0160021.767.31.6[93]
Cu/ZnO–Pd/CNT/
HZSM-5
2505.025,00018.950.30.58[104]
Pd–Cu–Al2O3/HZSM-52602.0300011.353.60.33[105]
1 Selectivity of DME in the products without CO. 2 STYDME is the space–time yield of DME.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, C.; Liu, Z. Perspective on CO2 Hydrogenation for Dimethyl Ether Economy. Catalysts 2022, 12, 1375. https://doi.org/10.3390/catal12111375

AMA Style

Liu C, Liu Z. Perspective on CO2 Hydrogenation for Dimethyl Ether Economy. Catalysts. 2022; 12(11):1375. https://doi.org/10.3390/catal12111375

Chicago/Turabian Style

Liu, Chang, and Zhongwen Liu. 2022. "Perspective on CO2 Hydrogenation for Dimethyl Ether Economy" Catalysts 12, no. 11: 1375. https://doi.org/10.3390/catal12111375

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop